Fluids 04 00081 - Pub
Fluids 04 00081 - Pub
Fluids 04 00081 - Pub
Article
Turbulence Model Assessment in Compressible Flows
around Complex Geometries with Unstructured Grids
Guillermo Araya
High Performance Computing and Visualization Laboratory, Department of Mechanical Engineering,
University of Puerto Rico, Mayaguez, PR 00681, USA; araya@mailaps.org
Received: 23 February 2019; Accepted: 15 April 2019; Published: 28 April 2019
Abstract: One of the key factors in simulating realistic wall-bounded flows at high Reynolds numbers
is the selection of an appropriate turbulence model for the steady Reynolds Averaged Navier–Stokes
equations (RANS) equations. In this investigation, the performance of several turbulence models
was explored for the simulation of steady, compressible, turbulent flow on complex geometries
(concave and convex surface curvatures) and unstructured grids. The turbulence models considered
were the Spalart–Allmaras model, the Wilcox k-ω model and the Menter shear stress transport (SST)
model. The FLITE3D flow solver was employed, which utilizes a stabilized finite volume method
with discontinuity capturing. A numerical benchmarking of the different models was performed
for classical Computational Fluid Dynamic (CFD) cases, such as supersonic flow over an isothermal
flat plate, transonic flow over the RAE2822 airfoil, the ONERA M6 wing and a generic F15 aircraft
configuration. Validation was performed by means of available experimental data from the literature
as well as high spatial/temporal resolution Direct Numerical Simulation (DNS). For attached or mildly
separated flows, the performance of all turbulence models was consistent. However, the contrary was
observed in separated flows with recirculation zones. Particularly, the Menter SST model showed the
best compromise between accurately describing the physics of the flow and numerical stability.
Keywords: Navier–Stokes equations (RANS); Direct Numerical Simulation (DNS); turbulence models;
compressible flow
1. Introduction
Wall-bounded turbulent flows at high Reynolds numbers are mainly characterized by a wide range
of time and length scales (Pope [1]). In Direct Numerical Simulation (DNS), the full Navier–Stokes
(NS) equations are generally solved to capture all time scales and eddies in the flow. This is
an accurate approach that supplies extensive information, but it demands the use of a very fine
mesh, and significant computational resources, to properly capture features in the order of the
Kolmogorov scales [2]. Filtered Navier–Stokes equations, with an additional sub-grid scale stress
term, are employed in Large Eddy Simulation (LES), in which large scales or eddies are directly
resolved and the effects of the small scales are modeled [3]. The LES approach still requires high
mesh resolution in boundary layer flows, particularly, in the near-wall region. Furthermore, the use
of hybrid approaches (RANS-LES) to overcome the fine resolution needed in wall-bounded flows is
an emerging and promising area [4]; nevertheless, there is still much ground to cover for industrial
applications. From the perspective of performing turbulent industrial flow simulations, the workhorse
is still the use of the Reynolds Averaged Navier–Stokes equations (RANS), which are obtained by time
averaging the full NS equations. In this approach, almost the full power spectra of velocity fluctuations
and turbulent kinetic energy are modeled, while capturing the very large turbulent scales of low
frequencies. These equations require a model or closure to compute the Reynolds stresses, which arise
from the convective terms of the NS equations after applying the time averaging process. The selection
2. Governing Equations
The unsteady compressible Navier–Stokes equations are expressed, over a 3D domain Ω ⊂ <3
with closed surface Γ, in the integral form
Z Z Z
∂Ui
dx + Fij n j dx = Gij n j dx, (1)
Ω ∂t ∂Ω ∂Ω
where ~n = (n1 , n2 , n3 ) is the unit normal vector to Γ. In addition, the unknown vector of conservative
variables is expressed as
ρ
Ui = ρui (2)
ρe
where ρ is the fluid density, ui denotes the ith component of the velocity vector and e is the specific
total energy and the inviscid and viscous flux vectors are expressed as
ρu j 0
Fij = ρui u j + pδij Gij = (3)
τij
ui (ρe + p) uk τkj − q j
is the deviatoric stress tensor, where µ is the dynamic viscosity and δij is the Kronecker delta.
The quantity q j = −k∂T/∂x j is the heat flux, where k is the thermal conductivity and T is the
absolute temperature. The viscosity varies with temperature according to Sutherlands’s law and the
Fluids 2019, 4, 81 3 of 24
Prandtl number is assumed to be constant and equal to 0.72. In addition, the medium is assumed to be
calorically perfect.
3. Solution Procedure
ΓK ΓK
∑ ∑
IJ IJ
Cj = AΓK n j I , Dj = AΓK n j I , (5)
I I
K ∈Γ I J K ∈Γ BIJ
ΓK
where AΓK is the area of facet ΓKI , n j I is the outward unit normal vector of the facet, Γ BIJ is the set of
I
ΓK
dual mesh faces on the computational boundary touching the edge between nodes I and J and n j I
denotes the normal of the facet in the outward direction of the computational domain. The numerical
integration of the fluxes over the dual mesh segment associated with an edge was performed by
assuming the flux to be constant, and equal to its approximated value at the midpoint of the edge,
i.e., a form of mid point quadrature. The calculation of a surface integral for the inviscid flux over the
control volume surface for node I is defined as follows,
IJ
Z Cj
∑ ∑
J IJ
F j n j dx ≈ F jI + F j + D j F jI , (6)
∂Ω I J ∈Λ I
2
J ∈Λ BI
where Λ I denotes the set of nodes connected to node I by an edge and Λ BI denotes the set of nodes
connected to node I by an edge on the computational boundary. Thus, the last term is non-zero only in
a boundary node. Similar formula can be implemented for the viscous fluxes. The use of edge based
data structure has become widely used due to its efficiency in terms of memory and CPU requirements,
compared to the traditional element based data structure, especially in three dimensions.
Fluids 2019, 4, 81 4 of 24
The resulting discretization is basically central difference in character. Therefore, the addition of
a stabilizing dissipation is required for practical flow simulations. This was achieved by replacing the
physical flux function by a consistent numerical flux function, such as the Jameson–Schmidt–Turkel
(JST) flux function [13] or the Harten–Lax–van Leer-Contact (HLLC) solver [14]. Discontinuity
capturing may be accomplished by the use of an additional harmonic term in regions of high pressure
gradients, identified by using a pressure switch.
4. Solver Parallelization
A parallel implementation of the flow solver was achieved by using the single program multiple
data concept in conjunction with standard Message Passing Interface (MPI) routines for message
passing. In the parallel implementation, at the start of each time step, the interface nodes of the
sub-domain with the lower number obtained contributions from the interface edges. These partially
updated nodal contributions were then broadcasted to the corresponding interface nodes in the
neighboring higher-number domains. A loop over the interior edges was followed by the receipt of
the interface nodal contributions and the subsequent updating of all nodal values. The procedure
was completed by sending the updated interface nodal values from the higher domain back to the
corresponding interface nodes of the lower domain. In addition, the computation and communication
processes took place concurrently. The employed approach was a global procedure in which the
agglomeration was performed across the parallel domain boundaries. Global agglomeration ensures
equivalence with a sequential approach for any number of domain, but increases the communication
load in the inner–grid mapping [8,16].
00 00
τijR = −ρui u j , (7)
which is the Favre averaged Reynolds stress tensor. The most straightforward approach is to associate
the unknown Reynolds stresses with the computed mean flow quantities by means of a turbulence
Fluids 2019, 4, 81 5 of 24
model or closure. If the Boussinesq hypothesis is applied, this results in a linear relationship to the
mean flow strain tensor through the eddy viscosity µt [17],
!
∂Ui ∂Uj 2 ∂Uk 2
τijR = µt + − δ − ρkδij , (8)
∂x j ∂xi 3 ∂xk ij 3
where k is the turbulent kinetic energy. The eddy viscosity depends on the velocity and the length scales
of the turbulent eddies, i.e., µt ∼ k1/2 `, where ` is the turbulence length scale. In this study, one- and
two-transport-equation models were considered, in which additional partial differential equations
were solved to describe the transport of the eddy viscosity. Consequently, nonlocal and history
effects on µt were taken into account. In the one-equation turbulence model of Spalart–Allmaras [18],
a combination of the turbulent scales, through the viscosity-like variable, ν̄, is obtained by solving
an empirical transport equation. Two-equation turbulence models are complete, because transport
equations are solved for both turbulent scales, i.e., the velocity and the length scale. The original
k-ω model [17] exhibits a freestream dependency of ω, which is generally not present in the k-e
model. Menter [19] combined the advantages of both models by means of blending functions,
which permits the switching from k-ω, close to a wall, to k-e, when approaching the edge of a boundary
layer. A further improvement by Menter [19] is a modification to the eddy viscosity, based on the
idea of the Johnson–King model, which establishes that the transport of the main turbulent shear
stresses is crucial in the simulations of strong Adverse Pressure Gradient (APG) flows. This new
approach is called the Menter shear–stress transport model (SST). In particular, the Menter SST
turbulence model is well-known for its good performance in boundary layer flows subjected to APG,
with eventual separation.
χ3
µt = ρν̃ , (10)
χ3 + c3v1
where ν is the molecular kinematic viscosity of the fluid, χ = ν̃/ν, and cv1 is equal to 7.1.
Fluids 2019, 4, 81 6 of 24
Additionally, St∞ = U∞ t/L and Re∞ = ρU∞ L/µ∞ are the Strouhal and Reynolds numbers,
respectively; Uj represents the time Favre-averaged velocity; and µ∞ is the freestream dynamic
viscosity. The rest of the parameters and constants are defined as follows:
ν̃
S = |ω | + f v2 , (11)
Re∞ κ 2 d2
χ
f v2 = 1 − , (12)
1 + χ f v1
!1
1 + c6w3 6
fω = g , (13)
g6 + c6w3
g = r + cw2 r r5 − 1 , (14)
ν̃
r = min , 10 , (15)
Re∞ Sκ 2 d2
f t2 = ct3 exp(−ct4 χ2 ), (16)
ωt2 2
2 2
f t1 = ct1 gt exp −ct2 d + gt d t , (17)
(∆U )2
∆U
gt = min 0.1, , (18)
ωt ∆x
where |ω | is the vorticity magnitude, d is the distance from a given point to the nearest wall, cb1 = 0.1355,
σ = 2/3, cb2 = 0.622, κ = 0.41, cω1 = cb1 /κ 2 + (1 + cb2 )/σ, cω2 = 0.3, cω3 = 2, ct1 = 1, ct2 = 2, ct3 = 1.1,
and ct4 = 2. Moreover, the variable ∆U is the difference in velocity between the point and the
associated trigger point, dt indicates the closest distance to such a trigger point or curve, ωt is the
vorticity magnitude at the associated trigger point, and ∆x is the surface grid spacing at this point.
The convective term is conveniently rewritten in conservative form,
which introduces an extra source term. Nevertheless, the contribution of this source term is
negligible [20]; thus, it is ignored in the present formulation.
Furthermore, the corresponding normalized transport equations in the Wilcox k-ω model [17]
for the turbulent kinetic energy, k, and the specific dissipation rate, ω, in compressible flows read
as follows:
" #
∂(ρUj k)
1 ∂(ρk) ∂Uj 1 ∂ µt ∂k
+ = τij − β k ρωk + µ+ (20)
St Re∞ ∂x j
| ∞ {z } | {zj }
∂t ∂x ∂x j | {z } σk ∂x j
| {z } dissipation term | {z }
transient term conv. term production term diffusion term
" #
1 ∂(ρω ) ∂(ρUj ω )
ω ∂Uj 1 ∂ µt ∂ω
+ = α τij − β ω ρω 2 + µ+ , (21)
St∞ ∂t ∂x j k ∂x j | {z } Re∞ ∂x j σω ∂x j
dissipation term
| {z } | {z } | {z } | {z }
transient term conv. term production term diffusion term
where β k = 0.09, β ω = 3/40, σk = 2, σω = 2 and α = 5/9. The normalized eddy viscosity is defined as:
In the Menter SST model for the ω equation, an extra term is considered in Equation (21), which
is called the cross-diffusion term,
ρσω2 ∂k ∂ω
2(1 − F1 ) , (23)
ω ∂x j ∂x j
| {z }
cross-diffusion term
and !
2ρσω2 ∂k ∂ω
CDkω = max , 10−10 . (25)
ω ∂x j ∂x j
Thus, the blending function F1 generates values close to one far from the wall (k-e model) and
almost zero values near the edge of the boundary layer (k-ω model). More details can be found in [19].
The main differences of Menter SST equations with respect to the Wilcox k-ω equations [17] can be
summarized as follows: (i) the consideration of a cross-diffusion term in the ω equation (k-e model),
which makes the model insensitive to the freestream boundary condition of ω; (ii) the implementation
of a stress limiter for the maximum value of ω as well as a production limiter to impede the build-up
of turbulence in stagnation zones; and (iii) the application of a blending function to compute the
corresponding constants of the k-e and k-ω models.
where tu stands for turbulence unknowns (ν̃, k and ω). Furthermore, the volume integrals were
calculated using the midpoint rule and the gradients appearing in the model were calculated as follows:
Z Z
∂Ui
dx = Ui n j dx, (27)
ΩI ∂x j ∂Ω I
where VI is the volume of the control volume. Equations (27) and (28) are expressed only for the
velocity but can be applied to any flow parameter. The second-order diffusion term was calculated
using the compact stencil form according to Equation (3.38) in [16], where the gradient along the edges
are evaluated by means of the compact finite difference scheme.
Fluids 2019, 4, 81 8 of 24
6νw
ωo = , (29)
βy2w Re L
where νw is the laminar kinematic viscosity at the wall, β is a constant (= 3/40), yw is the local first
off-wall point and Re L is the Reynolds number.
where Γe is the inlet engine surface and nej is the corresponding unit normal. Furthermore,
the independent variables of the turbulence model equations were prescribed fixed values at engine
inlets and outlets. In general, freestream values for the turbulence variables were set at the engine
inflow and outflow.
Fluids 2019, 4, 81 9 of 24
shown in Figure 1c) from the Menter SST model. The starting and ending points of the flat plate were
represented by two cross-sectional YZ cutting planes. The full X −length of the flat plate was about
87δre f . Figure 1c presents the natural evolution of k∗ in the direction of the flow, which was mainly
responsible for triggering turbulence.
Figure 2 shows the streamwise variation of the skin friction coefficient, C f , as a function of
Rex . Generally speaking, the corresponding values obtained by Spalart–Allmaras, Wilcox k-ω and
Menter SST turbulence models depicted an excellent agreement with theoretical correlations from
White [22] and Schlichting [23], particularly by the end of the flat plate. However, the Menter SST
model exhibited a shorter transition and the C f profile quickly tended to realistic fully-turbulent
values downstream from the leading edge. Additionally, Direct Numerical Simulation (DNS) data
from Araya and Jansen [24] are also included for a supersonic isothermal flat plate (Tw /T∞ = 2.25) at
a freestream Mach number, M∞ , of 2.5 and low Reynolds numbers. Again, the Menter SST turbulence
model exhibited a good agreement of C f with high spatial/temporal numerical results. Generally
speaking, the mean streamwise velocity along the boundary layer followed the DNS profile of Araya
and Jansen [24] and the 1/7 power-law distribution in the outer region (i.e., y/δ > 0.1), as shown in
Figure 3a for Rex = 3.4 × 107 . Nevertheless, all velocity profiles computed from turbulence models
depicted some deviations in the inner region (where the 1/7 power law did not work properly)
from the DNS profile. The thermal boundary layer plays a crucial role in the transport phenomena
of compressible wall-bounded flows; consequently, its accurate understanding and modeling are
very important. In other words, the coupled behavior of the velocity and thermal field should be
properly represented. In Figure 3b, the computed temperature distributions, T/T∞ , as a function
of the mean streamwise velocity, U/U∞ , are plotted at a streamwise station where Rex = 3.4 × 107 .
For comparison, the theoretical Crocco–Busemann relation (see page 502 in White [22]) is also included.
Fluids 2019, 4, 81 10 of 24
The Crocco–Busemann relation assumes a linear variation of the total enthalpy across the boundary
layer in zero pressure gradient with a unitary turbulent Prandtl number. Furthermore, present RANS
thermal profiles showed good agreement with the quadratic Crocco–Busemann relationship. It is
worth mentioning that the Crocco–Busemann relation is a function of the local wall-temperature;
therefore, a slightly different theoretical profile was obtained for each turbulence model.
(a) (b)
(c)
Figure 1. Mesh schematic (a); close-up of the near wall region (b); and iso-contours of turbulent kinetic
energy from Menter SST (c) in the supersonic flat plate.
(a) (b)
Figure 3. Mean streamwise velocity (a); and temperature distribution (b) in the supersonic flat plate.
In general, the three models are in good agreement with the experimental results and other
simulation values, as shown in Table 1. Furthermore, the pressure (C p ) and skin friction (C f ) coefficients
are depicted in Figure 4, together with the corresponding experimental data from the AGARD AR–138
report [25]. The predicted values of C p in the upper surface looked very similar in all turbulence models.
The weak shock at x/c ≈ 0.05 was not fully captured by turbulence models with discrepancies in the
order of 9% in peaks of C p . Downstream, a quasi-zero pressure gradient zone with almost constant
values of C p and about half-chord in length was observed. The second and stronger shock located at
x/c ≈ 0.55 was accurately captured by all models. Perhaps, the Menter SST slightly separated from the
experimental values by the end of the shock (i.e., x/c ≈ 0.57). The strong shock was characterized by
Fluids 2019, 4, 81 12 of 24
a sharp increase of wall pressure. The presence of a very strong APG induced a small flow separation
zone, which is discussed below. Beyond this zone of “pockets” with supersonic flow, the wall pressure
kept increasing towards the trailing edge, but at a moderate rate. In addition, the numerical predictions
of C p on the lower surface were almost identical to the experimental values given by all turbulence
τw
models. The skin friction coefficient is defined as C f = 1/2ρ U2
, where τw is the wall shear stress, and ρe
e e
and Ue are the density and velocity at the edge of the boundary layer, respectively. From the results
in Figure 4b, it was inferred that the closest values to experiments were produced by the Menter
SST model in the upper side, particularly at the strong shock location (i.e., x/c ≈ 0.55). Furthermore,
the Spalart–Allmaras and Wilcox k-ω models were observed to perform similarly; however, the Wilcox
k-ω was the only turbulence model that predicted a shock-induced separation due to the presence of
a very strong APG, with a small recirculating zone around 0.55 < x/c < 0.60, as shown in Figure 4b.
All models slightly underpredicted the experimental value of the skin friction on the lower surface.
Figure 5 shows iso-contours of the Mach number for the Wilcox k-ω model. This transonic regime
was characterized by the formation of regions or “pockets" with supersonic flow. Since the flow must
return to its freestream conditions by the time of reaching the trailing edge, this caused the formation
of a strong shock by x/c ≈ 0.55 where subsonic flow was observed beyond that point.
During the mesh generation, a good resolution was ensured by clustering the first-off wall point
well inside the linear viscous layer (i.e., y+ < 4) to accurately predict the skin friction coefficient.
Accordingly, the distance distribution of the first off-wall point in wall units, ∆z+w , along the RAE2822
airfoil and based on the Menter SST model is shown in Figure 6 together with the airfoil coordinates
in the right vertical axis. It can be seen that ∆z+ w was around 0.4 in the vicinity of the leading edge
for the lower surface; however, ∆z+ w was lower than 0.1 in the upper surface, where the flow faced
more complex pressure changes. Consequently, the high spatial resolution of the employed mesh was
demonstrated. In Figure 7, profiles of the mean streamwise velocity downstream of the strong shock in
the upper surface (i.e., at x/c = 0.65 and 0.75) are depicted. Notice that the mean streamwise velocity
was normalized by the local velocity at the edge of the boundary layer, Ue . In general, for velocity
profiles shown in Figure 7a,b, the Menter SST model produced the most accurate velocity profiles
of the three turbulence models tested in this study, when compared to experimental values from
Cook et al. [25]. It is interesting to point out that both velocity profiles at x/c = 0.65 and 0.75 were
located in a zone of very strong APG or increasing pressure, as shown in Figure 4a. Furthermore,
the Shear Stress Transport (SST) formulation by Menter [19] focused on improving the performance of
this turbulence model in adverse pressure gradient flows. Thus, the results depicted in Figure 7a,b
support the previous statement. In addition, the corresponding displacement thickness, δ∗ , momentum
thickness, θ ∗ , and, shape factor H (= δ∗ /θ ∗ ) were computed at these stations. These integral boundary
layer parameters (i.e., δ∗ and θ ∗ ) are defined as follows for compressible flows,
Z δ
∗ ρU
δ = 1− dz, (31)
0 ρe Ue
U
Z δ
∗ ρU
θ = 1− dz, (32)
0 ρe Ue Ue
where subscripts e stands for values at the edge of the boundary layer. In Table 2, the computed
values of δ∗ , θ ∗ and H are exhibited for the three turbulence models together with experimental
data from Cook et al. [25]. Furthermore, it was established that the three models analyzed in this
investigation gave a similar level of accuracy on the calculation of the boundary layer parameters;
nevertheless, the Menter SST was slightly superior to the other models with an average discrepancy
of 5% with respect to experiments. In addition, the knowledge of the shape factor, H, in turbulent
flows is important in assessing how strongly the APG is imposed. In fact, the higher is the value of
H, the stronger is the APG. Similarly, it is well known that any flow subjected to strong deceleration
is prone to separation, where most of the turbulence models find enormous difficulties. In this
Fluids 2019, 4, 81 13 of 24
opportunity, we picked up two stations in the RAE2822 airfoil (i.e., x/c = 0.65 and 0.75) with strong
APG or high shape factors. The idea was to evaluate the performance of the turbulence models in
extreme conditions. Generally speaking, the Menter SST model showed the best performance in
describing the physics of the flow subjected to strong APG.
Table 2. Boundary layer parameters at x/c = 0.65 and 0.75 in the RAE2822 airfoil.
(a) (b)
Figure 4. Pressure (a); and skin friction (b) coefficient in the RAE2822 airfoil.
Figure 6. Distance distribution of the first off-wall point in wall units along the RAE airfoil for the
Menter SST model.
in Figure 8. All three models captured reasonably well the corresponding location of both shocks at
each wing span section, given by abrupt modifications on C p . The most significant discrepancies were
observed at y/(b/2) = 0.8, where none of the three models were able to appropriately represent the
shock within 0.25 < x/c < 0.35. However, a similar behavior of the Spalart–Allmaras and Wilcox
k-ω turbulence models at y/(b/2) = 0.8 was reported by Huang et al. [31], Jakirlić et al. [32] and
Nielsen and Anderson [30]. The skin friction coefficient C f is plotted in Figure 9 at spanwise stations
y/(b/2) = 0.9 and 0.95. These figures show the good agreement among all turbulence models about
the location of the flow separation point x/c ∼ 0.25 and 0.2 for y/(b/2) = 0.9 and 0.95, respectively.
The downstream recovery of C f was different in all turbulence models, with the Wilcox k-ω and Menter
SST turbulence models predicting similar reattachment lengths and Spalart–Allmaras producing the
largest reattachment length. The Menter SST model induced higher values for the skin friction
downstream of the reattachment point, which was consistent with the largest values of CD f riction
shown in Table 3.
Figure 8. Cont.
Fluids 2019, 4, 81 16 of 24
Recirculation zone
Figure 10. Iso-surfaces of streamwise velocity over the upper surface in the ONERA M6 wing at
α = 3.06◦ (flow from left to right).
(a)
(b)
Figure 11. Iso-contours of streamwise velocity in Menter SST at y/(b/2) = 0.9 (a) and zoom over the
leading edge (b) in the ONERA M6 wing at α = 3.06◦ (flow from right to left).
Fluids 2019, 4, 81 18 of 24
In Figure 13a, iso-contours of the pressure coefficient C p over the entire F15 aircraft configuration
are depicted. Strong shock waves were observed on the canopy over the cock pit, leading edge of
wings and between the vertical tails. The cut plane in Figure 13b at x/L x ≈ 0.67 depicted high values
of C p toward the wing tip. In addition, a top view of iso-contours of C p are shown in Figure 14.
The value of the pressure coefficient at four points (labeled as I–IV in Figure 14) were compared with
flight and wind tunnel data from Webb et al. [33], as shown in Figure 15. In general, the present
numerical results are in good agreement with the experimental wind tunnel data. Furthermore, it can
be noticed that the pressure coefficients given by the Menter SST model showed a better match with
experimental data than those calculated by Spalart–Allmaras and Wilcox k-ω models. In addition,
the most significant discrepancies of the numerical results of C p with experiments were found in the
region 0.3 < x/L x < 0.5, just above the turbine intake. The discrepancy can be attributed to the
difference between the estimated engine mass flow prescribed and the used engine mass flow during
the flight test. This was emphasized by the significant scattering between the experimental data (in
flight and wind tunnel), particularly, at Points I and II. In this zone, the different turbulence models
predicted different separation regions, which are discussed below. Hence, this resulted in different
local flow patterns that induced different pressure coefficients.
Fluids 2019, 4, 81 19 of 24
(a) Front view of the whole grid configuration. (b) Cross-sectional cut at x/L x ≈ 0.67.
(a) Whole F15 aircraft configuration. (b) Cross-sectional cut at x/L x ≈ 0.67.
Figure 13. Iso-contours of the pressure coefficient in the F15 aircraft configuration (C p range of color
contour as in Figure 14).
y/Ly
1
0.5
I II III IV
0 0.5 1 x/Lx
Figure 14. Iso-contours of pressure coefficient in the upper surface of the F15 aircraft.
Fluids 2019, 4, 81 20 of 24
0.2
II
III
0
I
-0.2
IV
-0.4
Cp
-0.6
-1.2
0.3 0.4 0.5 0.6 0.7 0.8 0.9
x/Lx
Figure 15. Variation of the pressure coefficient in the top surface along x/L x of F15 aircraft for present
numerical data (open symbols) experimental data (closed symbols).
The skin friction coefficient C f at a spanwise location of y = 1m or y/Ly = 0.15, where Ly = 6.52 m
is the semi-span length of the aircraft, is plotted in Figure 16. The study of the drag coefficient due
to friction showed that this section had the largest discrepancy between the three turbulence models,
as observed in Figure 16a. Figure 16b indicates that the Wilcox k-ω model predicted two separation
bubbles at x/L x ≈ 0.325 and 0.36, respectively, while the other two models predicted a single separated
region (i.e., at x/L x ≈ 0.35) with the Menter SST predicting a much larger separation region than
that computed by the Spalart–Allmaras model. Furthermore, the iso-contours of streamwise velocity
(normalized by the freestream velocity U∞ ) by the Menter SST model, as shown in Figure 17, indicated
the presence of a lengthy recirculation zone with negative values of the velocity above the turbine
intake, given by the shock wave-boundary layer interaction.
(a) (b)
Figure 16. Skin friction coefficient in the upper surface of F15 at y/Ly = 0.15 (a); and zoom of the
separated flow zone (b).
Fluids 2019, 4, 81 21 of 24
Figure 17. Iso-contours of the streamwise velocity at y/Ly = 0.15 (flow from right to left).
Figures 18 and 19 show the pressure coefficients on the upper and lower surfaces of the wing
at two different spanwise locations, y/Ly = 0.36 and 0.59, respectively. In addition, the vertical
coordinates, z, of the local profile are represented by the vertical axis on the right. It is shown in
Figure 18a that, at y/Ly = 0.36, the C p profiles predicted by the three turbulence models showed
similar upper peak values in the vicinity of the leading edge, i.e., C p ≈ −1. Hence, it can be concluded
that the pressure difference between the surface static pressure to the freestream static pressure was
equal to the freestream dynamic pressure in absolute value at this location. On the other hand,
the Wilcox k-ω model predicted a stronger second shock wave located at x/L x ≈ 0.7. In the lower side,
a moderate APG zone was observed in 0.55 < x/L x < 0.6, as shown in Figure 18b, followed by a nearly
ZPG region. Furthermore, Figure 19a shows similar peak values of C p in the upper surface and close to
the leading edge at station y/Ly = 0.59 by all three turbulence models. However, the Spalart–Allmaras
model predicted a much stronger second shock located further downstream of the leading edge (at
x/L x ≈ 0.7) than those predicted by the Menter SST and Wilcox k-ω model. All turbulence models
exhibited similar C p distributions in the lower surface, as shown in Figure 19b.
(a) (b)
Figure 18. Pressure coefficients of F15 in the upper (a) and lower (b) surfaces at y/Ly = 0.35.
Fluids 2019, 4, 81 22 of 24
(a) (b)
Figure 19. Pressure coefficients of F15 in the upper (a) and lower (b) surfaces at y/Ly = 0.59.
8. Conclusions
An assessment of three popular turbulent models was performed in 3D aerodynamic cases.
The FLITE3D flow solver, in conjugation with the Spalart–Allmaras, the Wilcox k-ω and the Menter
SST turbulence models, was applied to the supersonic flat plate, RAE2822 airfoil, ONERA M6 wing,
and F15 aircraft.
The Menter SST model exhibited a short developing section in the supesonic flat plate downstream
of the edge. The transonic RAE2822 airfoil possessed a strong shock–boundary layer interaction
with an induced separation. The ONERA M6 wing exhibited a very complex flow phenomena
such as double shocks, strong APG and streamline curvature effects. In addition, the computed
global lift and drag coefficients (CL and CD ) over the F15 aircraft by all turbulence models were very
consistent; however, local pressure coefficients by the Menter SST model were in better agreement
with experimental values in [33].
Generally speaking, numerical results have been very consistent for the three turbulence models.
For attached flows or middle separated flows, there was not a clear superiority of the two-equation
turbulence models over the one-equation Spalart–Allmaras model. However, the contrary was
observed in separated flows with recirculation zones. Particularly, the Menter SST model showed
the best compromise between accurately describing the physics of the flow and numerical stability.
The Shear Stress Transport (SST) formulation might be the reason for this. Future investigation will
focus on unsteady flow simulations. This could be more appropriate in realistically capturing the
complex turbulent structures governed by shock-boundary layer interactions and highly-separated
flows; in particular, at the transonic regime.
Funding: This material was based upon work supported by the Air Force Office of Scientific Research (AFOSR)
under award number FA9550-17-1-0051.
Acknowledgments: The author acknowledges Oubay Hassan and Kenneth Morgan for valuable insight. Christian
Lagares is acknowledged for important discussion.
Conflicts of Interest: The author declare no conflict of interest.
References
1. Pope, S.B. Turbulent Flows; Cambridge University Press: Cambridge, UK, August 2000.
2. Moin, P.; Mahesh, K. Direct Numerical Simulation: A Tool in Turbulence Research. Annu. Rev. Fluid Mech.
1998, 30, 539–578. [CrossRef]
Fluids 2019, 4, 81 23 of 24
3. Meneveau, C.; Katz, J. Scale-invariance and turbulence models for large-eddy simulation. Annu. Rev.
Fluid Mech. 2000, 32, 1–32. [CrossRef]
4. Frohlich, J.; von Terzi, D. Hybrid LES/RANS methods for the simulation of turbulent flows. Prog. Aerosp. Sci.
2008, 44, 349–377. [CrossRef]
5. Catalano, P.; Amato, M. An evaluation of RANS turbulence modelling for aerodynamic applications.
Aerosp. Sci. Technol. 2003, 7, 493–509. [CrossRef]
6. Knight, D.; Yan, H.; Panaras, A.G.; Zheltovodov, A. Advances in CFD prediction of shockwave turbulent
boundary layer interactions. Prog. Aerosp. Sci. 2003, 39, 121–184. [CrossRef]
7. Manzari, M.T.; Hassan, O.; Morgan, K.; Weatherill, N.P. Turbulent flow computations on 3D unstructured
grids. Finite Elem. Anal. Des. 1998, 30, 353–363. [CrossRef]
8. Sørensen, K.A.; Hassan, O.; Morgan, K.; Weatherill, N.P. A multigrid accelerated hybrid unstructured mesh
method for 3D compressible turbulent flow. Comput. Mech. 2003, 31, 101–114. [CrossRef]
9. Peiró, J.; Peraire, J.; Morgan, K. FELISA System Reference Manual. Part 1—Basic Theory; Swansea Report
C/R/821/94; University of Wales: Wales, UK, 1994.
10. Peraire, J.; Morgan, K.; Peiró, J. Unstructured finite element mesh generation and adaptive procedures for
CFD. In Applications of Mesh Generation to Complex 3D Configurations; AGARD: Paris, France, 1990.
11. Hassan, O.; Morgan, K.; Probert, E.J.; Peraire, J. Unstructured tetrahedral mesh generation for
three-dimensional viscous flows. Int. J. Numer. Methods Eng. 1996, 39, 549–567. [CrossRef]
12. Weatherill, N.P.; Hassan, O. Efficient three–dimensional Delaunay triangulation with automatic boundary
point creation and imposed boundary constraints. Int. J. Numer. Methods Eng. 1994, 37, 2005–2039. [CrossRef]
13. Jameson, A.; Schmidt, W.; Turkel, E. Numerical simulation of the Euler equations by finite volume
methods using Runge–Kutta timestepping schemes. In Proceedings of the 14th Fluid and Plasma Dynamics
Conference, Palo Alto, CA, USA, 13–25 June 1981; AIAA Paper; pp. 81–1259.
14. Harten, A.; Lax, P.D.; van Leer, B. On upstreaming differencing and Godunov-type schemes for hyperbolic
conservation laws. SIAM Rev. 1983, 25, 35. [CrossRef]
15. Morgan, K.; Peraire, J.; Peiro, J.; Hassan, O. The computation of 3-dimensional flows using unstructured
grids. Comput. Methods Appl. Mech. Eng. 1991, 87, 335–352. [CrossRef]
16. Sørensen, K.A. A Multigrid Accelerated Procedure for the Solution of Compressible Fluid Flows on
Unstructured Hybrid Meshes. Ph.D. Thesis, University of Wales, Swansea, Wales, 2002.
17. Wilcox, D.C. Turbulence Modeling for CFD; DWC Industries, Inc.: La Canada, CA, USA, 2006.
18. Spalart, P.R.; Allmaras, S.R. A one equation turbulence model for aerodynamics flows. In Proceedings of the
30th Aerospace Science Meeting and Exhibit, Reno, NV, USA, 6–9 January 1992; AIAA Paper 92-0439.
19. Menter, F.R. Review of the shear-stress transport turbulence model experience from an industrial perspective.
Int. J. Comput. Fluid Dyn. 2009, 23, 305–316. [CrossRef]
20. Geuzaine, P. An Implicit Upwind Finite Volume Method for Compressible Turbulent Flows on Unstructured
Meshes. Ph.D. Thesis, Université de Liège: Liège, Belgium, 1999.
21. Spalart, P.R.; Rumsey, C.L. Effective Inflow Conditions for Turbulence Models in Aerodynamic Calculations.
AIAA J. 2007, 45, 2544–2553. [CrossRef]
22. White, F.M. Viscous Fluid Flow; McGraw Hill: New York, NY, USA, 1974.
23. Schlichting, H. Boundary Layer Theory, McGraw Hill: New York, NY, USA, 1968.
24. Araya, G.; Jansen, K. Compressibility effect on spatially-developing turbulent boundary layers via DNS.
In Proceedings of the 4th Thermal and Fluids Engineering Conference (TFEC2019), Las Vegas, NV, USA,
14–17 April 2019.
25. Cook, P.; McDonald, M.; Firmin, M. Aerofoil RAE2822 Pressure Distributions and Boundary Layer and Wake
Measurements; Report AR-138; AGARD: Paris, France, 1979.
26. Hirschel, E.H. Finite Approximations in Fluid Mechanics II: DFG Priority Research Program, Results 1986–1988;
Notes on Numerical Fluid Mechanics; Vieweg: Braunschweig, Germany; Wiesbaden, Germany, 1989;
Volume 25.
27. Swanson, R.C.; Rossow, C.C. An efficient solver for the RANS equations and a one-equation turbulence
model. Comput. Fluids 2011, 42, 13–25. [CrossRef]
28. Schmitt, V.; Charpin, F. Pressure Distributions of the ONERA M6 Wing at Transonic Mach Numbers;
Report AR-138; AGARD: Paris, France, 1979.
Fluids 2019, 4, 81 24 of 24
29. LeMoigne, A.; Qin, N. Variable-fidelity aerodynamic optimization for turbulent flows using a discrete adjoint
formulation. AIAA J. 2004, 42, 1281–1192.
30. Nielsen, E.J.; Anderson, W.K. Recent improvements in aerodynamic sesign optimization on unstructured
meshes. AIAA J. 2002, 40, 1155–1163. [CrossRef]
31. Huang, J.C.; Lin, H.; Yang, J.Y. Implicit preconditioned WENO scheme for steady viscous flow computation.
J. Comput. Phys. 2009, 228, 420–438. [CrossRef]
32. Jakirlić, S.; Eisfeld, B.; Jester-Zurker, R.; Kroll, N. Near-wall, Reynolds-stress model calculations of transonic
flow configurations relevant to aircraft aerodynamics. Int. J. Heat Fluid Flow 2007, 28, 602–615. [CrossRef]
33. Webb, L.; Varda, D.; Whitmore, S. Flight and Wind-Tunnel Comparisons of the Inlet/Airframe Interaction of the
F-15 Airplane; Technical Paper 2374; NASA: Washington, DC, USA, 1984.
c 2019 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).